Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Environmental Research 186 (2020) 109616

Contents lists available at ScienceDirect

Environmental Research
journal homepage: www.elsevier.com/locate/envres

Biogasoline production from linoleic acid via catalytic cracking over nickel T
and copper-doped ZSM-5 catalysts
Haswin Kaur Gurdeep Singha, Suzana Yusupa,∗, Armando T. Quitainb,c,∗∗, Bawadi Abdullaha,
Mariam Ameena, Mitsuru Sasakib, Tetsuya Kidab, Kin Wai Cheaha
a
Biomass Processing Lab, Center of Biofuel and Biochemical Research, Institute of Self-Sustainable Building, Universiti Teknologi PETRONAS, 32610, Bandar Seri
Iskandar, Perak, Malaysia
b
Faculty of Advanced Science and Technology, Kumamoto University, 2-39-1 Kurokami, Chuo-ku, Kumamoto, 860-8555, Japan
c
International Research Organization for Advanced Science and Technology, Kumamoto University, 2-39-1 Kurokami, Chuo-ku, Kumamoto, 860-8555, Japan

ARTICLE INFO ABSTRACT

Keywords: Catalytic cracking of vegetable oil mainly processed over zeolites, and among all the zeolites particularly HZMS-
Biogasoline 5 has been investigated on wide range for renewable and clean gasoline production from various plant oils.
Catalytic cracking Despite the fact that HZSM-5 offers a higher conversion degree and boost aromatics yield, the isomerate yield
Linoleic acid reduces due to high cracking activity and shape selectivity of HZSM-5. Hence, to overcome these problems, in
Isoparaffin
this study the transition metals, such as nickel and copper doped over HZSM-5 were tested for its efficiencies to
Aromatics
improve the isoparaffin compounds. The catalysts were screened with linoleic acid in a catalytic cracking re-
action conducted at 450 ᵒC for 90 min in an atmospheric condition in batch reactor. Then, the gasoline com-
position of the organic liquid product (OLP) was analysed in terms of paraffin, isoparaffin, olefin, naphthenes
and aromatics (PIONA). The results showed that Cu/ZSM-5 produced the highest liquid yield of 79.1%, at the
same time reduced the production of gas and coke to 18.8% and 0.7%. Furthermore, the desired isoparaffin
composition in biogasoline increased from 1.6% to 6.8% and at the same time reduced the oxygenated and
aromatic compounds to 15.4% and 59.7%, respectively. The linoleic acid as model compound of rubber seed oil,
in the catalytic cracking reaction provides a clearer understanding of the process. Besides, the water gas shift
(WGS) reaction in catalytic cracking reaction provides insitu hydrogen production to saturate the branched
olefin into the desired isoparaffin and the aromatics into naphthenes.

1. Introduction bioethanol has been widely used in Brazil replacing the conventional
gasoline to fuel vehicles. Though bioethanol has higher octane number,
Recently, The World Energy Outlook Report 2019 (BP, 2019) stated evaporation temperature, burning limit and speed, the main limitations
that the global energy consumption increased rapidly by 2.9% in year are low energy density, high chances of corrosion, low combustion
2018. As a result, the carbon emissions also increased at the highest rate luminescence and greater tendency to form azeotropic mixture with
by 2% in seven years. This was primarily attributed by the increase in water. Furthermore, a different combustion engine is required for cars
the number of passenger cars reflecting the transportation sector. The running on bioethanol because the fuel is made of bioalcohols not hy-
rising population and social development are expected to significantly drocarbons, resulting in lack of confidence in bioethanol as the new
boost the energy consumption in future. Additionally, the diminishing emerging fuel to substitute the conventional gasoline. Therefore, the
petroleum reserves will be insufficient to meet the increasing energy better alternative would be biogasoline due to its similar hydrocarbon
demand of the coming generation, if alternative solutions are not properties to the conventional gasoline and thus can be used directly in
considered today. internal combustion engine without engine modification. Furthermore,
Renewable energy from biomass is the most promising alternative it is proven to have a higher combustion energy and is more fuel effi-
solution because biomass derived fuels are categorised as clean and eco- cient compared to ethanol.
friendly fuels (Sher et al., 2020). The first generation biofuel such as The selection of technological pathway to produce biogasoline is


Corresponding author.
∗∗
Corresponding author. Faculty of Advanced Science and Technology, Kumamoto University, 2-39-1 Kurokami, Chuo-ku, Kumamoto, 860-8555, Japan.
E-mail addresses: drsuzana_yusuf@utp.edu.my (S. Yusup), quitain@kumamoto-u.ac.jp (A.T. Quitain).

https://doi.org/10.1016/j.envres.2020.109616
Received 31 January 2020; Received in revised form 19 April 2020; Accepted 27 April 2020
Available online 30 April 2020
0013-9351/ © 2020 Elsevier Inc. All rights reserved.
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

dependent on the physical state of the feedstock employed. Liquid product (Taufiqurrahmi and Bhatia, 2011). This was due to its excellent
feedstock is the most favourable because it produces biogasoline di- deoxygenation tendency and shape selectivity. The HZSM-5 catalyst
rectly either by catalytic cracking, hydrocracking or thermal cracking constitutes of a 3-dimensional intersecting structure with pore size si-
(Ameen et al., 2017). While, the solid feedstock has to be first converted milar to aromatic compounds like toluene, xylene and benzene (Hu
to syngas via gasification process, bio oil via pyrolysis or liquefaction et al., 2017). For that reason, it promotes the formation of aromatics
process and ethanol via enzyme or acid hydrolysis process prior to compounds in gasoline fraction.
gasoline production by Fischer Tropsch and catalytic upgrading pro- Although aromatics boost the fuel's octane number, stringent caps
cess. Nevertheless, catalytic cracking is the most preferred route be- are imposed by the legislative body on aromatics in a fuel. This is be-
cause the reactions proceed at a much lower temperature to produce a cause fuel rich in aromatics escalates the released of unburnt hydro-
higher octane gasoline (Ong and Bhatia, 2010), no hydrogen supply is carbons (carbon monoxide) and aromatics (benzene) to the environ-
required, and the reaction can take place at atmospheric condition. ment. The carbon monoxide gas causes global warming while benzene
Hence, this method is considered economical with respect to energy is a proven toxic substance. The carcinogenic property of aromatics
consumption (Taufiqurrahmi and Bhatia, 2011). Several experimental could impair human's health and lead to fatality. Therefore, to over-
setups using fixed bed, fluidized bed and entrained flow reactor have come this, the benign isoparaffins compounds are augmented to com-
been used for heterogeneous catalytic cracking reaction. The fluidised pensate for the drop in octane quality due to the reduction of the aro-
bed reactor is mostly used in petroleum refineries due to short contact matics content in fuel (Boot et al., 2017). Besides, further
time with catalyst and thus reduces the coke deposition and enhances polymerization of the aromatic compounds causes coke formation re-
the gasoline production. The catalytic reaction using this reactor is sulting in catalyst deactivation during the cracking reaction.
known as fluid catalytic cracking (FCC) reaction (Chew and Bhatia, Therefore, several published papers stress the use of metal loading
2008; Ong and Bhatia, 2010). especially transition metals on ZSM-5 to enhance the active acid sites of
The vegetable oils are commonly used for catalytic cracking to the catalyst owing to their beneficial bifunctional characteristics (acid
produce biogasoline. Both edible oils for example sunflower, rapeseed, and metal function). A study by Vichapund et al. (Vichaphund et al.,
and palm oil as well as non-edible oils such as jatropha and rubber seed 2015) presents the effect of various metals in HZSM-5 such as Ni, Mo,
oil (RSO) have been investigated. However, non-edible oils have shown Ga, Co and Pd on the deoxygenation and hydrogen transfer activity of
high potential based on researchers’ investigation due to food versus the catalyst. Ni and Co favour olefin formation while Mo, Ga and Pd
fuel conflicts (Hilten et al., 2011). The rubber seed oil extracted from promote the undesired aromatics formation via dehydrogenation. The
the seeds of the rubber tree is a potential non-edible oil source for deoxygenating trend reported was Ni > Pd > Co > Mo > Ga.
biogasoline production. It contains a higher degree of polyunsaturated Hence, nickel enhances deoxygenation reaction and reduces aromatics.
fatty acids, particularly linoleic acid which is very suitable for gasoline Besides, Zhao et al. (2015a), reported that using Zn/ZSM-5 enhanced
production. The reason is because the unsaturated fatty acid decom- the decarboxylation reaction by producing the highest CO2 yield that
poses and deoxygenates more rapidly than saturated fatty acids. Lino- favoured the removal of oxygen. Recently, Ahmad et al. (2016),
leic acid is a polyunsaturated fatty acid with two carbon-carbon double claimed that heterometallic nano oxides integrated with ZSM-5
bonds with a chemical formula of C18H32O2. As a result, RSO is more (Fe–Zn–Cu-ZSM-5 (31) gave highest gasoline yield by catalytic cracking
suitable for gasoline production than palm oil because the linoleic of palm oil.
composition of RSO (39%) is significantly higher than palm oil (9%) Though several studies have discussed on various metal modified
(Taufiqurrahmi and Bhatia, 2011). Likewise, previous work reported by ZSM-5 in response to biogasoline, aromatics, coke and oxygenates yield,
Bielansky et al. (2012) concluded that oleic acid (unsaturated fatty the study still lacks in experimenting modified ZSM-5 to increase the
acid) gave a higher gasoline yield compared to palmitic acid (saturated isoparaffin yield in the catalytic cracking reaction. The isomerization
fatty acid) in the presence of commercial FCC equilibrium catalyst. and hydrogenation reaction promote isoparaffin formation. Hydrogen
Catalyst affects the gasoline yield and composition in the catalytic transfer can also occur without direct hydrogen supply in a catalytic
cracking reaction. Several researchers have studied various types of cracking reaction. This is possible by exploiting the right catalyst that
catalyst in the catalytic cracking reaction to produce biogasoline as would initiate the hydrogen-transfer reaction. Furthermore, hydrogen
shown in Table 1. Among them, zeolite HZSM-5 showed higher gasoline could also be generated in-situ via water gas shift (WGS) reaction by
selectivity as it increased the aromatic content in the organic liquid using active WGS catalyst. Based on the reported literature, nickel

Table 1
Literatures on the catalytic cracking of vegetable oils for biogasoline production.
Catalyst Feedstock Reactor Yield (%) Reference

HZSM-5 Palm oil Fixed bed 48.5 Twaiq et al. (2004)


HZSM-5/Al2O3 Palm oil Fixed bed 47 Bhatia et al. (2009)
E-cat Cottonseed oil Fluidized bed 33.7 Li et al. (2009)
nano-ZSM-5 Used palm oil Fixed bed 37.05 Taufiqurrahmi et al. (2011)
H-AlMCM-41 Palm oil Fixed bed 29.4 Suyanta and Izul (2012)
HZSM-5 Palm oil Fixed bed 28.87 Sirajudin et al. (2013)
USY/ZSM-5 Sunflower oil Fixed bed 43.5 Doronin et al. (2014)
HZSM-5, Jatropha oil Fixed bed 50 Ramya et al. (2015)
H2SO4 Rubber seed oil Batch 48.3 Rengga et al. (2015)
V2O5 Sunflower oil Batch 29.2 Yigezu and Muthukumar (2015)
FeZnCuZSM-5 Palm oil Batch 59 Ahmad et al. (2016)
ZSM-5 Soybean oil Fixed bed 32.21 Emori et al. (2017)
USY Jatropha oil Fluidized bed 38.61 Zheng et al. (2017)
ZSM-5 Rubber seed oil Batch 34.86 Hassan et al. (2017)
E-cat Sunflower oil Fixed bed ≈40 Shimada et al. (2018)
HZSM-5 Waste cooking oil Fixed bed 35.0 Vu and Armbruster (2018)
HY Palm oil Fixed bed 34.25 Istadi et al. (2019)
ZSM-5 Palm oil Fixed bed 42.6 Makertihartha et al. (2020)

2
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

Fig. 1. Inconel batch reactor diagram (Chan et al., 2015).

Product (wt . )
Liquid or gas or coke product yield (%) = × 100
Linoleic acid fed (wt . ) (1)

Area of gasoline or above C or oxygenates


OLP composition (%) = × 100
Total area of liquid
(2)

Area of PIONA compounds


Gasoline composition (%) = × 100
Area of gasoline (3)

catalyst promotes deoxygenation activity and produce a large amount 2.3. Catalyst characterisation
of H2 gas during the catalytic pyrolysis of biomass in the presence of
steam (Kantarelis et al., 2014). The insitu hydrogen generation in water The X-ray Diffraction (XRD) patterns used to study the structure of
gas shift (WGS) reaction, shifts the hydrogen limited environment to the the catalysts were obtained from Rigaku X-ray diffractometer. The XRD
carbon-limited environment. The findings of Schaidle et al. (2015), patterns, were assessed at a range of 5ᵒ to 80ᵒ. The average crystallite
proves that Cu-modified BEA zeolite could produce high-value gasoline size was measured using Scherer's equation (Ameen et al., 2019). The
range alkanes via hydrogen integration simultaneously reducing the functional groups were determined by Fourier Transform Infrared
aromatics yield. Spectroscopy (FTIR) using JASCO FTIR-4100. The wavelength was kept
Based on literature review, so far limited literature has reported on in the range of 4000 to 400 cm⁻ˡ. The field emission electron microscopy
the catalytic cracking of model compound of rubber seed oil that is (FESEM) analysis was performed to study the catalyst morphology
linoleic acid for biogasoline production. Studying the model compound using Zeiss Supra 55 VP instrument. The surface area and pore property
gives a better understanding of the process and reaction routes. were examined by nitrogen adsorption-desorption method (BET ana-
Therefore, this study aims to investigate the catalytic cracking of lino- lysis) using Surface Area Analyzer and Porosimetry System, Model:
leic acid using HZSM5 as support, monometallic nickel and copper Micromeritics ASAP 2020.
doped over HZSM-5 and pure copper metal as catalyst in order to in- The characterization of the acid sites present on the catalyst was
crease the isoparaffin compounds and decrease the aromatics and studied using Temperature-Programmed Desorption of ammonia (NH3-
oxygenates in the gasoline fraction. TPD) analysis. The apparatus used was Model: TPD/R/O 1110. The
sample was pretreated at 500 ᵒC for 1 h under helium flow of 20 cm3/
min. After that, it was set to cool before saturating it for 30 min under
2. Materials and methods NH3 flow. Later, the sample was purged with helium again to remove
any physiosorbed NH3 before conducting NH3 desorption up to 550 ᵒC
2.1. Materials with a heating rate of 10 ᵒC/min.

The zeolite ZSM-5 in its ammonium form was obtained from Alfa
Aesar, USA. The copper (II) chloride dehydrate (99.0%) and nickel (II) 2.4. Catalytic cracking of linoleic acid
chloride hexahydrate (98.0%) as metal precursors as well as the feed-
stock linoleic acid (98.0%) were procured from Wako Chemicals, In this study, catalytic cracking approach was used to produce ga-
Japan. soline range hydrocarbons. The experiment was designed based on the
optimum condition reported by Rengga et al. (2015) on the catalytic
cracking of crude RSO over H2SO4 in a batch reactor. The experiment
2.2. Catalyst preparation was conducted under an atmospheric condition in an 8.8 mL Inconel
batch reactor as presented in Fig. 1. First, 0.5 wt% of catalyst were
Initially, the ZSM-5 in ammonium form was calcined at 550 ᵒC for introduced into the 4 mL of linoleic acid present in the Inconel batch
6 h forming HZSM-5. Later, 5 wt% of nickel and copper in HZSM-5 were reactor. Then, the electric furnace temperature was set to 450 ᵒC. Once
prepared using wet impregnation method (Vichaphund et al., 2015) the temperature stabilised, the Inconel batch reactor was placed into
using copper (II) chloride dihydrate and nickel (II) chloride hexahy- the electric furnace and the reaction was commenced for 90 min. As
drate solutions respectively. Subsequently, both the mixtures were soon as the reaction completes, the reactor was immediately condensed
stirred at 80ᵒC for 2 h. Then, the concentrated mixtures formed were in a bucket of water. After a while, centrifugation was carried out to
kept to dry overnight at 100ᵒC. The resulted Ni/ZSM-5 and Cu/ZSM-5 separate the solid from the liquid product. Then, once the liquid pro-
powders obtained later undergo calcination for 5 h at 540ᵒC in a fur- duct settles into 2 phase, the organic phase was segregated from the
nace. aqueous phase for further analysis using GCMS model HP6890 series

3
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

Fig. 2. XRD patterns of HZSM-5, Cu/ZSM-5 and Ni/ZSM-5.

with a HP-5MS capillary glass column (30 m long x 250 μm ID x


0.25 μm film thickness).
The compounds present in the organic liquid product (OLP) were
grouped into gasoline range hydrocarbons (C5–C12), above C12 hydro-
carbons and oxygenates. Further analysis of gasoline composition was
conducted based on PIONA grouping. The products distribution, OLP
composition and gasoline composition were calculated based on equa-
tions (1)–(3) respectively.

3. Results and discussion

3.1. Catalyst characterisation

3.1.1. XRD analysis


Fig. 2 illustrates the effect of incorporating metal in parent HZSM-5
on the catalyst structure. It was observed that metal addition did not
Fig. 3. FTIR spectra of HZSM-5, Cu/ZSM-5 and Ni/ZSM-5.
distort the XRD pattern of parent HZSM-5. Both metal deposited cata-
lysts still resemble their parent catalyst structure. Thus, preserving the
crystallinity of the catalyst. However, their intensities were reduced that the addition of metal affects the crystallinity of the parent catalyst.
with the addition of metal. The presence of copper was seen at peak Hence, irregularities in the catalysts morphologies were observed due
angles 43.1ᵒ, 50.2ᵒ and 73.7ᵒ while the presence of nickel was located at to sintering effect. Ahmad et al. (2016) also encountered similar effects.
peak angles 44.4ᵒ, 51.8ᵒ and 76.3ᵒ. The HZSM-5 characteristic peaks in Nevertheless, all the catalysts formed agglomerates because the crys-
all the catalysts were verified by the presence of peaks in the range of tallite size obtained from XRD presented in Table 2 was considerably
2θ = 7-10ᵒ and 23-25ᵒ (Bayat and Sadrameli, 2017). less than 1–2 μm (Miskolczi et al., 2019).
The irregularities of the crystals observed after metal loading on
3.1.2. FTIR analysis parent HZSM-5 changed the size and shape of the channels and con-
The FTIR spectrums displayed in Fig. 3 is consistent with the XRD sequently affected the gasoline selectivity and PIONA distribution for
pattern obtained. The metal impregnated catalyst showed similar trend the catalytic cracking of vegetable oils. Among the metal deposited
as compared to their parent HZSM-5 catalyst. Hence, the addition of catalysts, Cu/ZSM-5 showed less deterioration in crystals quality, which
metals did not significantly affect the functional groups present in the matched with the findings of Round et al. (2001). This was supported
parent catalyst but experienced slight decreased in transmittance in- by the dot mapping analysis which demonstrated that copper species
tensity. The peaks at approximate wavelengths of 3620 cm−1, was uniformly distributed compared to nickel species on the external
1640 cm−1, 1220 cm−1, 1100 cm−1, 800 cm−1, 550 cm−1 and surface of parent HZSM-5. Lesser dispersion was observed in Ni/ZSM-5
450 cm−1 represent acidic hydroxyl groups, water bending vibration, was primarily due to its bigger aggregate crystallite size in comparison
external asymmetric stretch, internal asymmetric stretch, external to Cu/ZSM-5.
symmetric stretch, double ring and internal tetrahedral bending vi-
bration respectively (Zhao et al., 2015a). 3.1.4. BET analysis
Fig. 5 illustrates the N2 physisorption curves of Cu/ZSM-5, Ni/ZSM-
3.1.3. FESEM analysis 5 and parent HZSM-5. It is apparent that the modified HZSM-5 and the
Fig. 4 shows the morphologies of parent HZSM-5 and the metal parent catalysts exhibit Type 1 isotherm, which signifies microporous
deposited HZSM-5 (copper and nickel) catalysts. The SEM image of solid. This is because the catalysts showed linear gas adsorption at a
HZSM-5 particles structure in Fig. 4a exhibited defined cubic crystal very low relative pressure forming a steep knee. After that, the gas
shape. This is consistent with the study reported by (Vichaphund et al., adsorption gradually decreases as relative pressure increases creating a
2015). However, in the other two images (Fig. 4b and c), it was noticed long horizontal plateau due to micropore filling (Wuamprakhon et al.,

4
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

Fig. 4. FESEM couple with dot mapping of (a) HZSM-5, (b) Cu/ZSM-5 and (c) Ni/ZSM-5.

2016). However, by introducing metals such as copper and nickel into Thus, it can be assumed that most of the metal would have been ad-
HZSM-5 the quantity of gas adsorbed reduced. Similarly, the specific sorbed only the external surface of HZSM-5 (Zhao et al., 2015b). On the
surface area also reduced with the metal impregnation nevertheless other hand, the average pore size increased slightly for both the metal
almost negligible change in the pore volume was observed in Table 2. modified HZSM-5 probably due to the blockage of metal species on the

5
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

Table 2
Surface area, porosity, aggregate crystallite size and acidity of HZSM-5, Ni/ZSM-5 and Cu/ZSM-5.
Catalyst Surface area (m2g−1) Pore Volume (cm3g−1) Pore size (nm) Aggregate crystallite size (nm) Acidity amount (μmol/g)

HZSM-5 432 0.262 2.43 21.5 6798.68


Ni/ZSM-5 400 0.264 2.63 26.5 7545.89
Cu/ZSM-5 391 0.258 2.63 19.5 7026.58

proton mobility in zeolites. On the other hand, the second peak ob-
served at high temperature region (400–600 ᵒC) signified the presence
of strong acid sites (Li et al., 2018). These peaks usually described
desorption of strongly bonded ammonia on Brønsted acid sites
(Rodríguez-González et al., 2007).
At low-temperature region parent HZSM-5 showed the highest
desorption peak however when metal like Cu and Ni were added the
desorption intensities reduced. In contrary, the metal-doped catalyst
exhibited higher desorption intensities than its parent catalyst at a
higher temperature region. Among them, Ni/ZSM-5 was the highest.
The acidity amount displayed in Table 2 further confirms the findings
obtained. Furthermore, the addition of metal also shifted the strong
acidic peaks to a higher temperature. The shifting was probably at-
tributed by the replacement of H+ with Cu2+ and Ni2+ respectively
which had affected the acidity of the catalyst. A similar trend was also
observed by Li et al. (2018).

3.2. Liquid, gas and coke distribution

Fig. 5. N2 adsorption-desorption isotherms of HZSM-5, Cu/ZSM-5 and Ni/ZSM- The product distribution of Ni/ZSM-5 and Cu/ZSM-5 were com-
5. pared to its parent HZSM-5 catalyst. The performance of pure copper
catalyst against copper-doped HZSM-5 was also tested. In terms of
product distribution as depicted in Fig. 7, it is observed that Ni/ZSM-5
micropores. Similar findings were reported by Zakaria et al. (2012) and
gave an opposite effect than Cu/ZSM-5 relative to its parent HZM-5.
Gulec et al. (Guleç et al., 2019).
Cu/ZSM-5 increased the liquid yield but reduced the gas and coke yield
while Ni/ZSM-5 showed the opposite trend. The similar result was
3.1.5. NH3TPD analysis
observed for Ni/ZSM-5 in comparison to ZSM-5 by Dewajani et al.
Fig. 6 shows NH3-TPD profiles of HZSM-5, Cu/ZSM-5 and Ni/ZSM-5
(2016) during the cracking of Nyamplung oil.
catalysts. The desorption peaks seen at two temperature regions at-
The product distributions were influenced by the catalyst properties
tributed to the presence of Brønsted and Lewis acid sites in the zeolite
because after metal impregnation with copper and nickel, the proper-
structure (Sandoval-Díaz et al., 2015). The peaks observed at low
ties of the parent HZSM-5 changed. The catalyst properties that affected
temperature region (100–300 ᵒC) usually represent weak Lewis acid
the catalytic cracking activity were mainly acidity and surface area (Li
sites of zeolites which correspond to desorption of weakly bonded
et al., 2018). Metal addition such as copper and nickel were reported to
ammonia. The concentration of this site has no significance effect on
increase the Lewis acid sites in the parent HZSM-5 (Li et al., 2018). This
the catalytic processes however, assumed to effectively facilitates the
type of acid sites reduces vigorous secondary cracking of liquid into
gaseous products unlike Brønsted acid sites. This explained the high

Fig. 6. NH3-TPD profiles of HZSM-5, Cu/ZSM-5 and Ni/ZSM-5. Fig. 7. Liquid, gas and coke yield of HZSM-5, Ni/ZSM-5, Cu/ZSM-5 and Cu.

6
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

liquid yield obtained for Cu/ZSM-5 but this was not true for Ni/ZSM-5,
though both had a high number acidic number sites. This is because, the
NH3TPD analysis for Ni/ZSM-5 showed that the catalyst had a higher
acidic strength than Cu/ZSM-5 due to the movement of the strong
acidic peaks to a much a higher temperature observed in Ni/ZSM-5
catalyst. As a result, the secondary cracking reactions occurred robustly
in Ni/ZSM-5 which led to lowest liquid yield.
In addition, other by-product such as coke is also produced during
the catalytic cracking reaction. A slightly higher coke yield was noticed
in Ni/ZSM-5 compared to HZSM-5 and Cu/ZSM-5. The increased in the
pore size together with the stronger Lewis acid sites found in the Ni/
ZSM-5 produced higher coke yield. This was because the increase in
pore size allowed higher accessibility of hydrocarbons to the pores
which resulted in polycondensation reactions that transpired the for- Fig. 9. Gasoline composition of HZSM-5, Ni/ZSM-5, Cu/ZSM-5 and Cu.
mation of heavy molecular weight hydrocarbons causing coke forma-
tion (Kantarelis et al., 2019). Furthermore, the presence of stronger hydrocarbons. This primary cracking is presumed to occur in the in-
Lewis acid sites acted as active sites to promote dehydrogenation re- ternal pores of the zeolites. The oxygenated compounds then undergo
action in the absence of hydrogen which resulted in the formation of deoxygenation to remove oxygen molecule via decarboxylation, dec-
coke precursor species on the catalyst surface (Triwahyono et al., arbonylation and dehydration reaction by emitting CO2, CO and H2O
2011). respectively to form heavy hydrocarbons. The pure heavy hydrocarbon
Therefore, the synergistic effect of copper with HZSM-5 resulted in chains then cracked to form lighter hydrocarbon chain known as par-
better performance on the product distributions compared to the stand- affin and olefins which can also be in a gaseous form during the sec-
alone copper catalyst, Ni/ZSM-5 and parent HZSM-5 catalyst. To fur- ondary cracking reaction. The lighter hydrocarbons could further un-
ther evaluate the effectiveness of these catalysts in promoting gasoline dergo isomerization, cyclization and aromatization in the presence of
range hydrocarbons formation, the liquid composition obtained for catalyst to form isoparaffins, naphthenes and aromatics respectively in
each catalyst was discussed in the following section. the gasoline range. Further polymerization of aromatics resulted in
coke formation. The condensation of triglycerides could also resulted in
3.3. Organic liquid product (OLP) distribution coke formation.
Nevertheless, gasoline range hydrocarbon represents the major
Fig. 8 illustrates the OLP compositions obtained for each catalyst constitute of the organic liquid products for all the catalyst with lesser
studied. It can be observed that gasoline range hydrocarbons (C5–C12) relative abundance of other by products such as heavy hydrocarbons
gradually increased with metal impregnation both for Ni/ZSM-5 and (above C12) and oxygenates. It is noted that Cu/ZSM-5 and Ni/ZSM-5
Cu/ZSM-5 relative to its parent HZSM-5 probably due to the increase in catalyst experienced slight increment in the relative abundance of
the number of active acid sites available on the catalyst (Ahmad et al., heavy hydrocarbons in their liquid composition. This is because tran-
2016). The presence of Brønsted acid sites as well as newly added Lewis sition metals like copper and nickel on HZSM-5 provides ideal en-
acid sites after metal modification of HZSM-5 promoted the cracking vironment for oligomerisation reaction of shorter chain alkene due to
reaction of long chain linoleic acid into gasoline range hydrocarbons the bifunctional characteristic of the metal sites and acid sites which
(Ahmad et al., 2016). However, the increment was not drastic even consequently results to heavier hydrocarbon formation (Vichaphund
after the addition of metal in HZSM-5. This explains that metal im- et al., 2015). However, between this two metal modified zeolites, Cu/
pregnation did not give a major impact on the gasoline distribution in ZSM-5 showed the least affinity for oligomerisation.
OLP but influenced the PIONA compounds distribution in gasoline as Interestingly, the addition of copper and nickel in HZSM-5 lowered
seen in Fig. 9. the relative abundances of oxygenates in the liquid composition in
Taufiqurrahmi & Bhatia (Taufiqurrahmi and Bhatia, 2011) ex- comparison to parent HZSM-5. This implies that metal modified HZSM-
plained the catalytic cracking mechanism of oils which resulted in the 5 shows a better deoxygenation tendency compared to the parent
organic liquid product distribution. The first step initiates the thermal HZSM-5. This concurs well with Vichapund et al. (Vichaphund et al.,
breakdown of heavy oxygenated hydrocarbons such as fatty acids that 2015) findings. This is because the increase in the Lewis acid sites on
are present in the vegetable oils to oxygenated hydrocarbons and heavy HZSM-5 with metal introduction enhances deoxygenation reaction of
oxygenated hydrocarbons (Ahmad et al., 2016). Among the catalysts
studied, Cu/ZSM-5 showed a superior performance in lowering the
oxygenated hydrocarbons while the stand-alone copper catalyst gave
the opposite results. This shows that the synergy effect of copper in
HZSM-5 improves the deoxygenating efficiency of the catalyst by low-
ering the relative abundance of oxygenates considerably from 34.7% to
15.4%. Larger relative abundance of oxygenates are undesirable as it
decreases the overall energy content of a fuel. As a result, for the same
driving cycle, a higher volumetric fuel is consumed (Martini et al.,
2013).
In conclusion, it confirms that Cu/ZSM-5 showed a highest deox-
ygenation tendency and gave the most gasoline range hydrocarbons as
compared to Ni/ZSM-5, parent HZSM-5 and copper catalyst. Hence, to
further evaluate the quality of the gasoline fraction, the discussion on
PIONA analysis were presented in the following section. Particularly,
relative abundances isoparaffin compounds were used as a basis to
Fig. 8. OLP composition of HZSM-5, Ni/ZSM-5, Cu/ZSM-5 and Cu. describe the most potential catalyst for producing high quality gasoline.

7
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

3.4. PIONA and oxygenates distribution promoted deoxygenation but the former gave a better result. Overall,
Cu/ZSM-5 showed superior performance in favouring gasoline-like
The PIONA distribution was determined by percentage area nor- hydrocarbons especially isoparaffin by lowering the undesired aro-
malization technique, based on peak areas obtained from the GC matics and oxygenates formation. This was an initial screening stage in
chromatograms. The remarkable correlation of copper in HZSM-5 that determining the best modified zeolite in favouring isoparaffin in gaso-
was worth noting in Fig. 9, was the selectivity towards isoparaffin line production from linoleic acid (model compound of RSO).
compounds, in comparison to its parent HZSM-5. The value increased
from 1.6% to 6.8%. However, both Ni/ZSM-5 and copper catalyst CRediT authorship contribution statement
showed low selectivity towards isoparaffin compounds. It has been
suggested by Schaidle et al. (2015) that incorporation of copper in Haswin Kaur Gurdeep Singh: Investigation, Writing - original
HZSM-5 produce H2 naturally. The author claimed that the H2 was draft. Suzana Yusup: Supervision, Conceptualization, Funding acqui-
produced via two ways. One is by the decomposition of dimethyl ether sition. Armando T. Quitain: Conceptualization, Resources. Bawadi
(DME) on metallic copper, present on zeolite's external surface to pro- Abdullah: Resources. Mariam Ameen: Writing - review & editing.
mote H2 activation. While, the second way is through alkane dehy- Mitsuru Sasaki: Resources. Tetsuya Kida: Resources. Kin Wai Cheah:
drogenation over Lewis acidic cationic copper sites, located within the Writing - review & editing.
zeolite pores to enhance hydrogen transfer reaction.
Furthermore, the CO produced from DME decomposition and water Declaration of competing interests
produced from DME homologation react in water-gas-shift (WGS) re-
action to form H2. Since copper is an active metal for WGS reaction, it The authors declare that they have no known competing financial
will enhance the H2 production. Therefore, employing the same prin- interests or personal relationships that could have appeared to influ-
ciple in catalytic cracking which gives the same by-products CO and ence the work reported in this paper.
H2O (deoxygenation reaction by-products) over metallic copper, could
also produce H2. Hence initiating hydrogenation even in the absence of Acknowledgements
H2 supply. This seems to be an innovative approach for saturating the
polyunsaturated feedstock like linoleic acid. Consequently favouring This research was financially supported by the Ministry of Higher
branched alkane (isoparaffin) formation derived from the hydrogena- Education (MOHE), Malaysia under the Long Term Research Grant
tion of C4 – C7 olefins through olefin catalytic cycle (Schaidle et al., Scheme (LRGS) and Higher Institution Centre of Excellence (HICoE).
2015). Alongside, slight increase in naphthenic and paraffinic com- The authors also acknowledge supports from the two collaborating
pounds were observed. universities - Universiti Teknologi PETRONAS (Malaysia) and
Ultimately, with the increased in the isoparaffin compounds, aro- Kumamoto University (Japan).
matics compounds decreased in Cu/ZSM-5 (59.7%) as compared to
both HZSM-5 (76.4%) and Ni/ZSM-5 (69.0%). This was probably due to References
decreased in the polycyclic aromatic hydrocarbons (PAH) formation,
produced via Diels-Alder reaction (Widayatno et al., 2016). Apart from Ahmad, M., Farhana, R., Raman, A.A.A., Bhargava, S.K., 2016. Synthesis and activity
that, the drop in the aromatics compounds might also be due to the evaluation of heterometallic nano oxides integrated ZSM-5 catalysts for palm oil
addition of metal in zeolite (Vichaphund et al., 2014). That is the cracking to produce biogasoline. Energy Convers. Manag. 119, 352–360. https://doi.
org/10.1016/j.enconman.2016.04.069.
reason that the aromatics compounds for Ni/ZSM-5 decreased even Ameen, M., Azizan, M.T., Yusup, S., Ramli, A., Yasir, M., 2017. Catalytic hydro-
though it did not increase the isoparaffin compounds. On the contrary, deoxygenation of triglycerides: an approach to clean diesel fuel production. Renew.
olefin compounds gradually increased in Cu/ZSM-5. This finding cor- Sustain. Energy Rev. 80, 1072–1088. https://doi.org/10.1016/j.rser.2017.05.268.
Ameen, M., Tazli, M., Ramli, A., Yusup, S., 2019. Catalytic hydrodeoxygenation of rubber
roborates with the explanation given by Schaidle et al. (2015). The seed oil over sonochemically synthesized Ni-Mo/γ -Al 2 O 3 catalyst for green diesel
author discovered that copper modified zeolite without cofed H2 production. Ultrason. Sonochem. 51, 90–102. https://doi.org/10.1016/j.ultsonch.
showed high selectivity for olefins by means of dehydrogenation reac- 2018.10.011.
BP, 2019. BP Statistical Review of World Energy 2019. (London).
tion. Bayat, A., Sadrameli, S.M., 2017. Production of renewable aromatic hydrocarbons via
In summary, the evidence from this PIONA analysis studies implies conversion of canola oil methyl ester (CME) over zinc promoted HZSM-5 catalysts.
that Cu/ZSM-5 is a potential catalyst for producing high quality bio- Renew. Energy 106, 62–67. https://doi.org/10.1016/j.renene.2017.01.014.
Bhatia, S., Mohamed, A.R., Shah, N.A.A., 2009. Composites as cracking catalysts in the
gasoline from the catalytic cracking reaction of linoleic acid which is a
production of biofuel from palm oil: deactivation studies. Chem. Eng. J. 155,
model compound of RSO. The considerable production of isoparaffins 347–354. https://doi.org/10.1016/j.cej.2009.07.020.
in the gasoline fraction over Cu/ZSM-5 catalyst justifies the improve- Bielansky, P., Weinert, A., Schonnberger, C., Reichhold, A., 2012. Gasoline and gaseous
ment of the gasoline quality. hydrocarbons from fatty acids via catalytic cracking. Biomass Convers. Biorefinery 2,
53–61. https://doi.org/10.1007/s13399-011-0027-x.
Boot, M.D., Tian, M., Hensen, E.J.M., Mani Sarathy, S., 2017. Impact of fuel molecular
4. Conclusions structure on auto-ignition behavior – design rules for future high performance ga-
solines. Prog. Energy Combust. Sci. 60, 1–25. https://doi.org/10.1016/j.pecs.2016.
12.001.
Catalytic cracking of linoleic acid was carried out in an Inconel Chan, Y.H., Yusup, S., Quitain, A.T., Tan, R.R., Sasaki, M., Lam, H.L., Uemura, Y., 2015.
batch reactor at 450 ᵒC for 90 min at atmospheric condition. The effect Effect of process parameters on hydrothermal liquefaction of oil palm biomass for
of HZSM-5, Cu/ZSM-5, Ni/ZSM-5 and copper catalyst in favouring ga- bio-oil production and its life cycle assessment. Energy Convers. Manag. 104,
180–188. https://doi.org/10.1016/j.enconman.2015.03.075.
soline type hydrocarbons in terms of PIONA were investigated in order Chew, T.L., Bhatia, S., 2008. Catalytic processes towards the production of biofuels in a
to evaluate the quality of the gasoline produced. Among the catalyst palm oil and oil palm biomass-based biorefinery. Bioresour. Technol. 99, 7911–7922.
studied, Cu/ZSM-5 gave the highest liquid yield of 79.1%, at the same https://doi.org/10.1016/j.biortech.2008.03.009.
Dewajani, H., Rochmadi, Purwono, S., Budiman, A., 2016. Effect of modification ZSM-5
time reduced the production of gas and coke to 18.8% and 0.7%.
catalyst in upgrading quality of organic liquid product derived from catalytic
Besides, both the Cu/ZSM-5 and Ni/ZSM-5 showed slight improvement cracking of Indonesian nyamplung oil (Calophyllum inophyllum). AIP Conf. Proc.
of gasoline relative content in the OLP in comparison to its parent 1755, 1–7. https://doi.org/10.1063/1.4958485.
Doronin, V.P., Potapenko, O.V., Lipin, P.V., Sorokina, T.P., 2014. Conversion of vegetable
HZSM-5. However, the former had a greater impact than the latter in
oils under conditions of catalytic cracking. Catalogue Index 6, 53–59. https://doi.
favouring gasoline range hydrocarbons. Furthermore, the synergism org/10.1134/S207005041401005X.
between copper and HZSM-5 promotes the desired isoparaffin produc- Emori, E.Y., Hirashima, F.H., Zandonai, C.H., Ortiz-Bravo, C.A., Fernandes-Machado,
tion. The value increased from 1.6% to 6.8%. In addition, both copper N.R.C., Olsen-Scaliante, M.H.N., 2017. Catalytic cracking of soybean oil using ZSM5
zeolite. Catal. Today 279, 168–176. https://doi.org/10.1016/j.cattod.2016.05.052.
and nickel modified HZSM-5 reduced the aromatics formation and

8
H.K. Gurdeep Singh, et al. Environmental Research 186 (2020) 109616

Guleç, F., Sher, F., Karaduman, A., 2019. Catalytic performance of Cu- and Zr-modified 10.1021/cs501876w.
beta zeolite catalysts in the methylation of 2-methylnaphthalene. Petrol. Sci. 16, Sher, F., Iqbal, S.Z., Liu, H., Imran, M., Snape, C.E., 2020. Thermal and kinetic analysis of
161–172. https://doi.org/10.1007/s12182-018-0278-2. diverse biomass fuels under different reaction environment : a way forward to re-
Hassan, S.N., Sarulnisah, R.F., Ahmed, M., Raman, A.A.A., Nik Sulaiman, N.M., 2017. newable energy sources. Energy Convers. Manag. 203, 112266. https://doi.org/10.
Biogasoline synthesis from rubber seed oil through thermal fluid catalytic cracking. 1016/j.enconman.2019.112266.
Aust. J. BASIC Appl. Sci. 11, 104–113. Shimada, I., Nakamura, Y., Ohta, H., Suzuki, K., Takatsuka, T., 2018. Co-processing of
Hilten, R., Speir, R., Kastner, J., Das, K.C., 2011. Production of aromatic green gasoline saturated and unsaturated triglycerides in catalytic cracking process for hydrocarbon
additives via catalytic pyrolysis of acidulated peanut oil soap stock. Bioresour. fuel production. J. Chem. Eng. Jpn. 51, 778–785. https://doi.org/10.1252/jcej.
Technol. 102, 8288–8294. https://doi.org/10.1016/j.biortech.2011.06.049. 17we187.
Hu, C., Xiao, R., Zhang, H., 2017. Ex-situ catalytic fast pyrolysis of biomass over HZSM-5 Sirajudin, N., Jusoff, K., Yani, S., Ifa, L., Roesyadi, A., 2013. Biofuel production from
in a two-stage fluidized-bed/fixed-bed combination reactor. Bioresour. Technol. catalytic cracking of palm oil. World Appl. Sci. J. 26, 67–71. https://doi.org/10.
https://doi.org/10.1016/j.biortech.2017.07.011. (in press). 5829/idosi.wasj.2013.26.nrrdsi.26012.
Istadi, I., Buchori, L., Anggoro, D.D., Riyanto, T., Indriana, A., Khotimah, C., Pangestu, Suyanta, S., Izul, F.I., 2012. Cracking of palm oil over H-AIMCM-41 catalyst. J. Chem.
F.A.S., 2019. Effects of ion exchange process on catalyst activity and plasma-assisted Chem. Eng. 6, 531–535.
reactor toward cracking of palm oil into biofuels. Bull. Chem. React. Eng. Catal. 14, Taufiqurrahmi, N., Bhatia, S., 2011. Catalytic cracking of edible and non-edible oils for
459–467. https://doi.org/10.9767/bcrec.14.2.4257.459-467. the production of biofuels. Energy Environ. Sci. 4, 1087. https://doi.org/10.1039/
Kantarelis, E., Yang, W., Blasiak, W., 2014. Effect of zeolite to binder ratio on product c0ee00460j.
yields and composition during catalytic steam pyrolysis of biomass over transition Taufiqurrahmi, N., Mohamed, A.R., Bhatia, S., 2011. Production of biofuel from waste
metal modified HZSM5. Fuel 122, 119–125. https://doi.org/10.1016/j.fuel.2013.12. cooking palm oil using nanocrystalline zeolite as catalyst: process optimization stu-
054. dies. Bioresour. Technol. 102, 10686–10694. https://doi.org/10.1016/j.biortech.
Kantarelis, E., Javed, R., Stefanidis, S., Psarras, A., Iliopoulou, E., Lappas, A., 2019. 2011.08.068.
Engineering the catalytic properties of HZSM5 by cobalt modification and post - Triwahyono, S., Jalil, A.A., Mukti, R.R., Musthofa, M., Razali, N.A.M., Aziz, M.A.A., 2011.
synthetic hierarchical porosity development. Top. Catal. 62, 773–785. https://doi. Hydrogen spillover behavior of Zn/HZSM-5 showing catalytically active protonic
org/10.1007/s11244-019-01179-w. acid sites in the isomerization of n -pentane. Appl. Catal. Gen. 407, 91–99. https://
Li, H., Shen, B., Kabalu, J.C., Nchare, M., 2009. Enhancing the production of biofuels from doi.org/10.1016/j.apcata.2011.08.027.
cottonseed oil by fixed-fluidized bed catalytic cracking. Renew. Energy 34, Twaiq, F., Mohamed, A.R., Bhatia, S., 2004. Catalytic cracking of palm oil into liquid
1033–1039. https://doi.org/10.1016/j.renene.2008.08.004. Fuels : kinetic study. Seventh Asia-Pasific Int. Symp. Combust. Energy Util. 1–8.
Li, C., Ma, J., Xiao, Z., Hector, S.B., Liu, R., Zuo, S., Xie, X., Zhang, A., Wu, H., Liu, Q., Vichaphund, S., Aht-ong, D., Sricharoenchaikul, V., Atong, D., 2014. Catalytic upgrading
2018. Catalytic cracking of Swida wilsoniana oil for hydrocarbon biofuel over Cu- pyrolysis vapors of Jatropha waste using metal promoted ZSM-5 catalysts: an ana-
modified ZSM-5 zeolite. Fuel 218, 59–66. https://doi.org/10.1016/j.fuel.2018.01. lytical PY-GC/MS. Renew. Energy 65, 70–77. https://doi.org/10.1016/j.renene.
026. 2013.07.016.
Makertihartha, I.G.B.N., Fitradi, R.B., Ramadhani, A.R., Laniwati, M., Subagjo, Muraza, Vichaphund, S., Aht-Ong, D., Sricharoenchaikul, V., Atong, D., 2015. Production of aro-
O., 2020. Biogasoline production from palm Oil : optimization of catalytic cracking matic compounds from catalytic fast pyrolysis of Jatropha residues using metal/
parameters. Arabian J. Sci. Eng. https://doi.org/10.1007/s13369-020-04354-4. HZSM-5 prepared by ion-exchange and impregnation methods. Renew. Energy 79,
Martini, G., Manfredi, U., Krasenbrink, A., Stradling, R., Zemroch, P.J., Rose, K.D., Hass, 28–37. https://doi.org/10.1016/j.renene.2014.10.013.
H., Maas, H., 2013. Effect of Oxygenates in Gasoline on Fuel Consumption and Vu, X.H., Armbruster, U., 2018. Catalytic cracking of triglycerides over micro/meso-
Emissions in Three Euro 4 Passenger Cars Contact Information. European porous zeolitic composites prepared from ZSM-5 precursors with varying aluminum
Unionhttps://doi.org/10.2790/1136. contents. React. Kinet. Mech. Catal. 1–14. https://doi.org/10.1007/s11144-018-
Miskolczi, N., Juzsakova, T., Soja, J., 2019. Preparation and application of metal loaded 1415-z.
ZSM-5 and y-zeolite catalysts for thermo-catalytic pyrolysis of real end of life vehicle Widayatno, W.B., Guan, G., Rizkiana, J., Yang, J., Hao, X., Tsutsumi, A., Abudula, A.,
plastics waste. J. Energy Inst. J. 92, 118–127. https://doi.org/10.1016/j.joei.2017. 2016. Upgrading of bio-oil from biomass pyrolysis over Cu-modified β-zeolite cata-
10.017. lyst with high selectivity and stability. Appl. Catal. B Environ. 186, 166–172. https://
Ong, Y.K., Bhatia, S., 2010. The current status and perspectives of biofuel production via doi.org/10.1016/j.apcatb.2016.01.006.
catalytic cracking of edible and non-edible oils. Energy 35, 111–119. https://doi.org/ Wuamprakhon, P., Wattanakit, C., Warakulwit, C., Yutthalekha, T., Wannapakdee, W.,
10.1016/j.energy.2009.09.001. Ittisanronnachai, S., Limtrakul, J., 2016. Direct synthesis of hierarchical ferrierite
Ramya, G., Sivakumar, T., Arif, M., Ahmed, Z., 2015. Application of microporous cata- nanosheet assemblies via an organosilane template approach and determination of
lysts in the production of biofuels from non edible vegetable oils and used restaurant their catalytic activity. Microporous Mesoporous Mater. 219, 1–9. https://doi.org/10.
oil. Energy sources, Part A recover. Util. Environ. Eff. 37, 878–885. https://doi.org/ 1016/j.micromeso.2015.07.022.
10.1080/15567036.2011.590855. Yigezu, Z.D., Muthukumar, K., 2015. Biofuel production by catalytic cracking of sun-
Rengga, W.D.P., Handayani, P.A., Kadarwati, S., Feinnudin, A., 2015. Kinetic study on flower oil using vanadium pentoxide. J. Anal. Appl. Pyrolysis 112, 341–347. https://
catalytic cracking of rubber seed (Hevea brasiliensis) oil to liquid fuels. Bull. Chem. doi.org/10.1016/j.jaap.2015.01.002.
React. Eng. Catal. 10, 50–60. https://doi.org/10.9767/bcrec.10.1.5852.50-60. Zakaria, Z.Y., Linnekoski, J., Amin, N.A.S., 2012. Catalyst screening for conversion of
Rodríguez-González, L., Hermes, F., Bertmer, M., Rodríguez-Castellón, E., Jiménez-López, glycerol to light olefins. Chem. Eng. J. 207– 208, 803–813. https://doi.org/10.1016/
A., Simon, U., 2007. The acid properties of H-ZSM-5 as studied by NH 3 -TPD and 27 j.cej.2012.07.072.
Al-MAS-NMR spectroscopy. Appl. Catal. Gen. 328, 174–182. https://doi.org/10. Zhao, X., Wei, L., Cheng, S., Huang, Y., Yu, Y., Julson, J., 2015a. Catalytic cracking of
1016/j.apcata.2007.06.003. camelina oil for hydrocarbon biofuel over ZSM-5-Zn catalyst. Fuel Process. Technol.
Round, C.I., Williams, C.D., Latham, K., Duke, C.V.A., 2001. Ni-ZSM-5 and Cu-ZSM-5 139, 117–126. https://doi.org/10.1016/j.fuproc.2015.07.033.
synthesized directly from aqueous fluoride gels. Chem. Mater. 13, 468–472. https:// Zhao, X., Wei, L., Cheng, S., Julson, J., 2015b. Optimization of catalytic cracking process
doi.org/10.1021/cm000591s. for upgrading camelina oil to hydrocarbon biofuel. Ind. Crop. Prod. 77, 516–526.
Sandoval-Díaz, Luís-Ernesto Gonzalez-Amaya, J.-A., Trujillo, C.-A., 2015. General aspects https://doi.org/10.1016/j.indcrop.2015.09.019.
of zeolite acidity characterization. Microporous Mesoporous Mater. 215, 229–243. Zheng, Q., Huo, L., Li, H., Mi, S., Li, X., Zhu, X., Deng, X., Shen, B., 2017. Exploring
https://doi.org/10.1016/j.micromeso.2015.04.038. structural features of USY zeolite in the catalytic cracking of Jatropha Curcas L. seed
Schaidle, J.A., Ruddy, D.A., Habas, S.E., Pan, M., Zhang, G., Miller, J.T., Hensley, J.E., oil towards higher gasoline/diesel yield and lower CO2emission. Fuel 202, 563–571.
2015. Conversion of dimethyl ether to 2,2,3-trimethylbutane over a Cu/BEA catalyst: https://doi.org/10.1016/j.fuel.2017.04.073.
role of Cu sites in hydrogen incorporation. ACS Catal. 5, 1794–1803. https://doi.org/

You might also like